[go: up one dir, main page]

0% found this document useful (0 votes)
34 views28 pages

Preview 2023 Electrochemistry Fundamentals Clarke

The document is a primer on electrochemistry, detailing its historical significance and fundamental principles, including thermodynamics and kinetics of electron transfer. It highlights the role of electrochemistry in daily life through applications like batteries and sensors, as well as its future potential in energy conversion and environmental solutions. The text aims to provide insights for both newcomers and experienced researchers in the field, emphasizing the importance of electrochemical interfaces.

Uploaded by

anmoldas8114
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
34 views28 pages

Preview 2023 Electrochemistry Fundamentals Clarke

The document is a primer on electrochemistry, detailing its historical significance and fundamental principles, including thermodynamics and kinetics of electron transfer. It highlights the role of electrochemistry in daily life through applications like batteries and sensors, as well as its future potential in energy conversion and environmental solutions. The text aims to provide insights for both newcomers and experienced researchers in the field, emphasizing the importance of electrochemical interfaces.

Uploaded by

anmoldas8114
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 28

https://pubs.acs.org/doi/book/10.1021/acsinfocus.

7e7020

This is a limited PDF preview of the primer. The entire work is available in
ePub3 and includes additional multimedia.

Electrochemistry Fundamentals

Thomas B. Clarke
Purdue University

Christophe Renault
Purdue University

Jeffrey E. Dick
Purdue University

Individual sales

Institutional sales
https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Preface

Since chemistry's infancy, various electrochemical phenomena have perplexed chemists, catalyzing important discoveries. For example, the world's
first widely publicized battery—Alessandro Volta's pile of zinc, silver, and brine-soaked pasteboard—begged important fundamental questions
ever since its disclosure in 1800. Within a few decades, several early elements were isolated using Volta's electrochemical pile (or devices that
improved upon his design), and the century-long debate over how the pile worked would eventually underscore the electric workings of chemistry.
Not long after, batteries improved, and the number of influential electrochemical experiments increased. James Joule's electrolysis experiments
helped demonstrate the equivalences between electrical, chemical, and heating effects that proved vital to the acceptance of the conservation
of energy and other thermodynamic laws. Svante Arrhenius' electrochemical studies of electrolyte solutions led to his discovery that ions are
charged species. Many of the fundamental principles of chemistry were either the direct consequence of or were informed by experiments in
electrochemistry. Throughout this text, we describe several of these experiments in more detail and show several examples of how electrochemistry
continues to be used to uncover new truths about nature today.

Perhaps you are less interested in the history of chemistry and electrochemistry and are more interested in how electrochemistry infiltrates
your daily life. Batteries and continuous glucose sensors you may use daily are devices based on electrochemical reactions. Electrochemistry is
also at the heart of several vital tools used to make discoveries in chemistry and other science labs today, as evidenced by pH sensors and gel
electrophoresis cells. Many of the devices of tomorrow will rely on knowledge of and discoveries in electrochemistry—fuel cells that efficiently
convert hydrogen fuel to usable energy; the carbon capture and conversion devices that will turn greenhouse gases into valuable products;
and the photoelectrochemical, semiconductor, and bioelectrochemical devices yet to come. The future will continue to harness and control
electrochemical reactions. All of these past, present, and future electrochemical contraptions and processes share a common feature: an interface
where charges are passed (i.e., an electrochemical interface). In this text, we cover many of the important features of these interfaces that will
prime you to begin diving into the exciting current research in this field.

While electrochemistry has helped answer fundamental questions about chemistry and proved helpful in our daily lives, fundamental questions
about electrochemistry have continued to arise, and not all have been answered. Within this text, we ask several of these fundamental questions:
What exactly happens when a metal electrode is submerged in a solution? What happens as current flows through an electrochemical cell? What
does the electrolyte environment look like near an electrode? What is important when designing a device that uses electricity to help controllably
catalyze a reaction or one that uses a chemical reaction to provide electricity? We present existing understandings to answer these deceptively
simple questions. We hope that offering some of the original experiments and results that helped answer these questions will help you better
understand the fundamentals of electrochemistry.

Whether you are a student starting electrochemistry research or a seasoned veteran in another scientific discipline wanting to gain a greater
appreciation and understanding of the complex phenomena at play in electrochemical devices, we hope our introduction to the field is both
informative and inspiring.

Thomas B. Clarke is a Ph.D. candidate in the Department of Chemistry at Purdue University. He received his B.S. in Chemistry
from the University of Notre Dame in 2018. While teaching high school chemistry and biology for two years upon graduation,
he earned his M.Ed. from the University of Notre Dame in 2020. He then began his graduate research in chemistry at
the University of North Carolina at Chapel Hill and has continued this work under the direction of Jeffrey Dick at Purdue
University, where he is using electrochemistry to study unique chemistry that happens in confined volumes and at liquid|
liquid interfaces. In addition to his research interests, he loves reading through the literature in chemistry education and
studying electrochemistry's role in the history of chemistry and science.

Christophe Renault received his Ph.D. in Molecular Electrochemistry from the University Paris 7 Denis-Diderot, Paris, in 2012.
After postdoctoral stays in the groups of Richard Crooks, Allen Bard, and Serge Lemay, he started his independent career
as a researcher at the Centre National de la Recherche Scientifique in France. He has worked since 2016 in the Laboratory
of Condensed Matter Physics at the Ecole Polytechnique in Palaiseau. He is interested in analytical and physical problems
related to electrochemistry. Dr. Renault is developing coupled opto-electrochemical measurements to decipher interactions
between electrified interfaces and individual molecules or particles. His expertise covers charge and mass transfer/transport
theories, coupled electrochemical and chemical reactions, numerical simulation, and instrumentation.

© 2023 American Chemical Society 1


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Jeffrey E. Dick received a B.S. in Chemistry from Ball State University (2013) and a Ph.D. in Chemistry from the University
of Texas at Austin (2017), where he studied under Prof. Allen J. Bard. In 2018, Jeffrey began his independent career at
the University of North Carolina at Chapel Hill and is now the Richard B. Wetherill Associate Professor of Chemistry at
Purdue University. Jeffrey's research group is broadly interested in fundamental and applied electrochemistry, believing that
developing new measurement tools can elucidate new truths of nature.

ACKNOWLEDGMENTS

J.E.D. would like to gratefully acknowledge the National Science Foundation (under grant CHE-2003587/2319925) and National
Science Foundation CAREER award (under grant CHE-2045672) for providing a platform to prioritize broader impacts, allowing
time to teach the world electrochemistry.

We would also like to thank the American Chemical Society for the opportunity to write this text. The entire team at ACS In
Focus, especially Sara Tenney, Maria Lokshin, Tricia Louvar, Benedicte Rossi, and the many individuals behind the scenes were
incredibly helpful throughout the process. They deserve great thanks.

Thomas B. Clarke, Christophe Renault, and Jeffrey E. Dick

DEDICATION

We dedicate this book to Prof. Allen J. Bard (University of Texas at Austin) on the occasion of his 90th birthday.

© 2023 American Chemical Society 2


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Brief Table of Contents

About the Series


Preface

1 Introduction to Electrochemistry—Past and Present

2 Thermodynamics of Electron Transfer at an Interface

3 Kinetics of Electron Transfer at an Interface

4 The Electrode | Electrolyte Interface

5 Inner-Sphere Reactions and


Electrocatalysis
Appendix A Derivation of the Nernst Equation

Appendix B Derivation of the Butler–Volmer and Current-Overpotential Equations


Bibliography
Glossary
Footnotes
Index

© 2023 American Chemical Society 3


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Detailed Table of Contents

About the Series


Preface

1 Introduction to Electrochemistry—Past and Present


1.1 Overview
1.2 The Origins of Electrochemistry—One Pile and a Century-Long Debate
1.3 Modern Applications to Industry and Research
1.4 That’s a Wrap

2 Thermodynamics of Electron Transfer at an Interface


2.1 Overview
2.2 Electromotive Force: Historical Insights and Modern Usage
2.3 Potentials
2.3.1 Potential Energy vs Potential
2.3.2 Potential Differences and Voltage
2.3.3 What Happens When a Metal Contacts a Solution?
2.3.4 Standard Reduction Potentials
2.3.5 Reference Electrodes
2.3.6 A Case Study: The Daniell Cell
2.3.7 Chemical and Electrochemical Potential
2.4 The Nernst Equation
2.4.1 Electrochemical Potential and Equilibrium
2.4.2 Derivation of the Nernst Equation
2.4.3 When the Nernst Equation Falls Short
2.5 Liquid Junction Potentials
2.5.1 What Happens When Two Electrolyte Solutions Come into Contact?
2.5.2 Ion Mobility, Conductivity, and Transference Numbers
2.5.3 How Problematic Are Liquid Junction Potentials in Potential Measurements?
2.5.4 Ways of Mitigating Liquid Junction Potentials
2.6 That’s a Wrap

3 Kinetics of Electron Transfer at an Interface


3.1 Overview
3.2 Homogeneous vs Heterogeneous Reaction Kinetics
3.3 A Model for Electrode Kinetics: The Butler–Volmer Equation
3.3.1 The Unique Features of Heterogeneous Electron-Transfer Reactions
3.3.2 Derivation of the Butler–Volmer Equation
3.3.3 Important Notes about the Butler–Volmer Equation
3.4 Practically Measuring Currents
3.5 Correlating Current Density with Overpotential: Tafel Plots
3.5.1 Tafel’s Observations
3.5.2 What Happens to the Potential when Current Flows?
3.5.3 Limitations of Tafel Plot Analysis

© 2023 American Chemical Society 4


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

3.6 A Model for Charge Transfer: Marcus Theory


3.6.1 Motivation and Foundations for Marcus Theory
3.6.2 Ramifications of Marcus Theory
3.6.3 Limitations of Marcus Theory
3.7 That’s a Wrap

4 The Electrode|Electrolyte Interface


4.1 Overview
4.2 Historical Insight: Arrhenius’s Nobel Prize
4.3 Electrolytes—What’s in Salt Water?
4.3.1 Conductivity and Conductance
4.3.2 Resistance and Resistivity
4.3.3 Kohlrausch’s Law
4.3.4 Nonaqueous Solutions
4.4 The Electrode|Solution Interface and the Electric Double Layer
4.4.1 Charging an Electrode
4.4.2 Mapping Potential in a Three-Electrode Cell
4.5 The Various Means of Mass Transfer
4.5.1 The Nernst–Planck Equation
4.5.2 The Diffusion Layer
4.5.3 Rotating Electrodes
4.5.4 Electrophoresis: Using Electric Fields to Move Charged Species
4.6 Electroanalytical Techniques
4.6.1 Potential Step or Amperometry
4.6.2 Linear Sweep Voltammetry and Cyclic Voltammetry
4.7 That’s a Wrap

5 Inner-Sphere Reactions and


Electrocatalysis
5.1 Overview
5.2 The Challenges in Driving Inner-Sphere Reactions
5.2.1 Potential Windows for Studying Electrochemical Reactions
5.2.2 Many Desired Electrochemical Reactions Proceed via Multielectron Pathways
5.2.3 The Roles of the Electrode and Adsorption in Inner-Sphere Reactions
5.2.4 Solvation Effects
5.2.5 Proton-Coupled Reactions
5.3 Considerations for Other Inner-Sphere Reactions
5.4 Features of Electrocatalysts: Materials and Design
5.5 Electrochemical Phase Transformations
5.5.1 Electrodeposition and Gas Evolution Reactions
5.5.2 Thermodynamics of Electrochemical Phase Transformations
5.6 That’s a Wrap
Appendix A Derivation of the Nernst Equation
Appendix B Derivation of the Butler–Volmer and Current-Overpotential Equations
B.1 The Butler–Volmer Equation
B.2 The Current-Overpotential Equation

© 2023 American Chemical Society 5


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Bibliography
Glossary
Footnotes
Index

© 2023 American Chemical Society 6


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

CHAPTER 1

Introduction to Electrochemistry—Past and


Present
1.1 Overview

1.2 The Origins of Electrochemistry—One Pile and a Century-Long Debate

1.3 Modern Applications to Industry and Research

1.4 That’s a Wrap

1.1 OVERVIEW

Electrochemistry is a discipline found at the intersection of electricity and chemistry, at the point of conversion between electrical potential and
chemical potential and, most often, at the interface between electron conductors and ionic conductors. It is the study of how electron-transfer
reactions occur at an electrode’s interface. By changing the potential of an electron conductor that is submerged in a solution, one can either
promote or prohibit electron-transfer reactions from occurring at the conductor’s interface with the solution. While a previous primer from
the ACS In Focus Series (1) has discussed the application of electrochemistry as a means of chemical analysis, this primer will dive into the
fundamentals of electron-transfer reactions. We will focus on the thermodynamics (CHAPTER 2) and kinetics (CHAPTER 3) of these reactions,
the effects of the electrode, electrolyte, and their interface (CHAPTER 4), and the mechanisms by which complex, multistep electron-transfer
reactions are facilitated by electrodes whose potential is modified (CHAPTER 5).

Fundamental electrochemistry underpins much of our daily lives, underlying the chemistry of batteries that power our devices, the processes that
corrode our materials, the methods by which we purify and treat water, and so much more. Before we discuss fundamental electrochemistry
principles that give rise to these and other phenomena, it is enlightening to discuss some of the history that got us here. Since the first
electrochemistry studies, key fundamental questions about the interplay between electrical and chemical phenomena have perplexed scientists,
even inciting decades-long debates. Take, for example, the world’s first electrochemical battery, the voltaic pile. This device was the first of its kind
to provide a sustained current source, but it preceded so many discoveries of electric and chemical phenomena that it is no wonder that it set off
a century-long debate concerning the source of the voltaic pile’s activity. Since this primer will touch on many of the same themes (e.g., relating
electric and chemical phenomena) and will reference historic experiments and discoveries, we take the time now to summarize the voltaic pile’s
discovery, the ensuing debate over the origins of its activity, and how it (and other electrochemical experiments) helped inform other discoveries
within chemistry.

THE ORIGINS OF ELECTROCHEMISTRY—ONE PILE AND A CENTURY-LONG


1.2
DEBATE
The story of electrochemistry arguably began at the turn of the 19th century when an Italian scientist assembled a pile of metals and soaked
pasteboard (cardboard). After being briefly fired from his professorship—as the French Revolutionary Wars waged on and control over his
university continued to switch hands—Alessandro Volta began working in a rudimentary lab that he set up in his private country house (2).
Without many resources or students, he continued his work on electricity, having already invented an early electrostatic generator (called the
electrophorus). He took notice of a set of experiments being performed in 1792 by another Italian scientist, Luigi Galvani, who found that when
dissimilar metals connected in a metallic arc were inserted into frog legs, the legs would begin to twitch (FIGURE 1.1a) (3). Galvani’s discovery
set forth theories about how electric phenomena may arise from biological origins. After all, electric eels had captured the attention of many
scientists of the era. Volta disagreed that the source of electric phenomena was biological and instead investigated ways to create similar effects
with inanimate materials (4).

© 2023 American Chemical Society 7


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

FIGURE 1.1 Two Early (c. 1790s) Electrochemical Experiments.

(a) Galvani’s frog leg experiment where a metal arc (consisting of two dissimilar metals connected at the center) was inserted into
frogs’ legs, which induced sudden twitching of the tissue. Reproduced from citation (5). (b) A model of the voltaic pile wherein pairs of
metal (e.g., zinc and copper) were stacked with brine-soaked pasteboard separating the pairs. The two ends of the pile could give quite
shocking sensations. Copied from citation (6).

With few resources and students, Volta assembled a stack of pairs of metals (i.e., zinc and copper or zinc and silver) that were in contact with
each other, but where the pairs were separated by a pasteboard soaked in brine (salt water) (7). To his surprise, this “pile”, as he called it (FIGURE
1.1b), was unlike any of the other electric contraptions of the time. The Leyden jar, for example, was an electrostatic device known at the time to
store electricity and severely shock people upon discharge. Benjamin Franklin was the first to use the term “battery” in 1747 to describe a series of
Leyden jars that he assembled that could be simultaneously charged or discharged (8). What made Volta’s pile so revolutionary was that it was a
constant source of electricity; the Leyden jars and other prior devices all needed to be recharged after a single discharge.

The voltaic pile caught the attention of many during this time. The timeline presented in FIGURE 1.2 shows the many experiments that followed
Volta’s discovery. After receiving Volta’s letter dated March 20, 1800, that described his pile’s construction and the sustained electrical activity it
could provide, the President of the Royal Society shared the results with Anthony Carlisle, who then shared it with William Nicholson (2). These
two immediately proceeded to perform experiments with voltaic piles and in early May 1800, were able to achieve the electrolysis of water by
connecting a voltaic pile to two platinum wires submerged in water (FIGURE 1.2). They instantly saw gases being generated from both wires and
rationalized that they had split water into its components: hydrogen and oxygen gas (9).

© 2023 American Chemical Society 8


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

FIGURE 1.2 Several Important Discoveries in the Understanding of Electrochemistry.

Volta disclosed his pile (March 20, 1800). Nicholson and Carlisle electrolyzed water a couple of months later (May 1800). Experiments
investigating oxygen’s role in the working of the pile were performed by Haldane and Davy (in October and November 1800,
respectively). Faraday introduced the term “ions” as neutral species transferring between electrodes (1834). Arrhenius proposed his
dissociation theory (1887). Nernst published his equation for the potential difference (E) between a metal and solution of metal ions
in terms of temperature (T), dissolution tension of the metal (P), and the osmotic pressure of the metal ions (p) (1889). Thomson
quantified the charge-to-mass ratio of cathode ray particles (1897); however, these were not identified as electrons until later. The term
“oxidation” came to mean the loss of electrons (1920s). Copied from Clarke et al. J. Chem. Educ. 2021 (10).

As the voltaic pile was being used to drive many reactions in experiments (besides water electrolysis, several elements would be isolated for the
first time via the pile’s electricity), early questions arose about how the voltaic pile worked. Other experiments in the same year examined the
electrical activity in different ambient environments. In October 1800, Henry Haldane placed voltaic piles in nitrogen-rich environments or early
vacuums and observed that their electric activity decreased, whereas in an oxygen-rich environment the activity was enhanced (11). One month
later, Sir Humphry Davy concluded that the electric activity of the voltaic pile was proportional to the degree of zinc oxidation occurring in the
pile (12). Note that at the time the term “oxidation” only meant “combination with oxygen” and was not used to describe loss of electrons for
over a century (see FIGURE 1.2) (13).

Nicholson and Davy saw the chemical implications of the voltaic pile and believed that chemical reactions occurring at the contact between the
metal and the brine solution were responsible for the electrical activity. However, this theory that chemistry gives rise to electric phenomena was
not immediately accepted—in fact the pile’s inventor, Volta, thought that it was simply the contact of the two dissimilar metals that caused a
force to push charges away from one another, a hypothesis that could explain both his pile and Galvani’s metallic arc experiment. Several other
scientists, collectively named “contact theorists,” agreed with Volta and presented experiments to defend this electricity by contact force theory.
Science historians have documented how this debate over a “contact” vs “chemical” explanation of the voltaic pile lasted for decades (14, 15).
This confusion and denial (by many) of the chemical explanation may sound preposterous to us now, but it is helpful to remember how little was
known at the time.

Many discoveries were prerequisite in developing our understanding of electrochemical cells. While different combinations of metals and
electrolyte solutions were used in the early 19th century to make batteries other than Volta’s, it was not until a series of discoveries that chemists
were able to more fully comprehend the workings of these batteries. The term ions, describing particles moving in solution, was proposed by
Michael Faraday (of the Faraday’s constant) in 1834 (16), and the fact that salts dissociated into charged species (our modern definition of ions)
was not known until Svante Arrhenius’s proclamation of such in 1887 (17). Nearly 90 years after Volta’s initial announcement, Wilhelm Nernst
was able to predict rather accurately the potential difference of an electrode and the electrolyte in terms of the temperature (T), dissolution tension

© 2023 American Chemical Society 9


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

of the metal (P) and the osmotic pressure of the metal ions (p) with his equation (FIGURE 1.2) (18). As we will see in SECTION 2.4, this is not
our modern way of expressing the common Nernst equation that is frequently used to predict the potential of an electrode or electrochemical cell.
Even Nernst did not have the full physical and chemical understanding that we do currently. This makes sense since J. J. Thomson would not
discover the charge-to-mass ratio of cathode ray particles (which would later be termed electrons) until 1897 (19), and, as mentioned previously,
the process of oxidation and reduction did not come to mean loss or gain of electrons until the 1920s (13). It is no wonder that 19th century
scientists struggled to explain the voltaic pile and debated the contact vs chemical theories of its workings: the necessary knowledge was missing.
While there was intense scrutiny and study of the voltaic pile’s workings in the early 19th century, it was not until very recently (2021) that a more
complete, modern description of the voltaic pile was established: zinc oxidation (loss of electrons) occurs at the anode and oxygen reduction (gain
of electrons) occurs at the cathode (10). This is just one example of how much there is still to learn—even from centuries-old science.

The importance of Volta’s and other electrochemical batteries in the development of broader chemistry knowledge is also inspiring. Volta’s pile
was directly used to isolate new elements. Nicholson and Carlisle were early users of the pile, isolating hydrogen and oxygen from the electrolysis
of water; others, like Sir Humphry Davy, used the pile for electrolysis to isolate elements such as sodium, potassium, and others (20). To Davy,
electrolysis was not just a new way to isolate elements, but a manifestation of the interconnectedness of chemical and electrical activity (21). This
served as a paradigm shift to him that matter was not inert (as Newton and others had thought), but was charged and was held together by
electrical forces (22).

By the mid-19th century, James Joule, after whom the SI unit of energy is named, performed many of his first experiments studying electricity
and electrochemical cells, where he found that the heat released by a wire was proportional to the square of current passing through the wire of
a given resistance over a specific period of time. Furthermore, he used electrolysis studies to establish the interconversion of electricity, chemistry,
and heat, even calculating the decomposition potential of water to rather accurate values (23). These early electrochemistry studies preceded and
likely informed Joule’s future work on mechanical equivalents in heat and the acceptance of the First Law of Thermodynamics.

In numerous other ways, electrochemical experiments have helped inform discoveries about various chemical phenomena. Answering questions
concerning fundamental electrochemical phenomena have frequently led to some of these insights. For example, as we will see in SECTION 4.3,
Arrhenius’ work with conductivity of electrolyte solutions led to his discoveries that salts and acids dissociate into charged species (which earned
him a Nobel Prize in chemistry and is why we have the Arrhenius definitions of acids and bases as species that produce proton or hydroxide
in solution). Electrochemistry has continued to be fundamental to uncovering new truths about nature, but, armed with more robust, modern
theories to explain the fundamentals, electrochemistry is additionally having a large impact in our daily lives.

1.3 MODERN APPLICATIONS TO INDUSTRY AND RESEARCH

Today, electrochemistry underpins our daily lives. Generally, the field can be divided into several subdisciplines, with much current attention
being paid toward energy applications as well as means of synthesis.

Of course, batteries have continued to develop since Volta’s initial version from 1800, but further improvements are reliant on a fundamental
understanding of electron-transfer reactions that occur at the two electrodes. The push toward renewable energy sources has also led to
electrochemistry as a viable alternative to producing fuels. Much current research is being conducted on electrolysis of water to produce hydrogen
fuel in the most efficient way. Fuel cells (electrochemical in nature) allow this hydrogen fuel to be turned back into electrical energy. Furthermore,
electrochemistry is considered the solution for reducing the amount of greenhouse gases, especially carbon dioxide, and converting these into
some valuable products and fuels.

Electrochemistry also provides an approach to controllably synthesize or deposit various chemicals and materials. Electroplating, for example, is a
common practice for depositing a thin layer of metal, alloy, or oxide onto a conductive surface. This can be decorative (as is the case when you
want to “redip” your white gold wedding ring) or practical (when you want to prevent corrosion of a material). Electrosynthesis, where electric
current can help drive and control synthetic reactions, has also been an exciting field for organic chemists. Much current research is focused on
finding inexpensive, corrosive-resistant electrode materials that can selectively electrocatalyze organic reactions at high yields (24).

Additionally, electrochemistry has been used for sensing and other analytical purposes. If an analyte of interest can participate in an oxidation or
reduction reaction, an electrochemical cell or device can be engineered to pass current to/from the analyte in solution to determine the analyte
concentration. Challenges arise when trying to electrochemically measure analytes that do not readily become oxidized or reduced. One common
research goal is finding another “reporter” species that is redox-active that can chemically react with the analyte or an enzyme which, when
catalyzing a reaction involving the analyte, produces a redox-active byproduct that can be measured electrochemically. Specific ligands or receptors
can also be employed to allow for the analyte/reporter reaction to occur near the electrode so that electrochemical responses can be measured.

Furthermore, other electrochemical sensors are based on measuring potentials to read out a concentration. A common example is the conventional
glass pH electrode in a pH meter, which acts as a glass ion-selective membrane that is sensitive to the relative concentrations of proton in the
test solution in which it is submerged. By measuring the potential across this glass membrane, the pH can be determined (25). Many other
ion-selective electrodes are used and continue to be developed for other sensing purposes (26).

© 2023 American Chemical Society 10


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Many other chemical parameters (besides concentrations) can also be measured with electrochemical cells. While several of these applications of
electrochemistry will be discussed as specific reactions in subsequent chapters, this text is aimed toward providing relevant information concerning
the fundamentals of electrochemical reactions.

1.4 THAT’S A WRAP

• The voltaic pile was the first publicized electrochemical battery, having been disclosed by Allessandro Volta in 1800. It was made of
stacks of zinc and silver (or copper) separated by pasteboard soaked in brine; however, improved alternative batteries quickly emerged
and have been continuing to be pursued ever since.

• The explanation for the voltaic pile’s ability to sustain electrical activity was the source of disagreement amongst scholars for about a
century, likely since several crucial physical and chemical concepts would not be understood for several more decades (e.g., the concept
of ions, Arrhenius’ electrolyte dissociation theory, the Nernst equation, the discovery of the electron, etc.).

• Current areas of research in electrochemistry include (but are not limited to) batteries, other forms of energy storage and conversion
(e.g., fuel cells, photovoltaics, supercapacitors, etc.) electrodeposition, electrosynthesis, electroanalytical methods, sensors, and more!

© 2023 American Chemical Society 11


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

APPENDIX A
Derivation of the Nernst Equation
We have already established the following definition of electrochemical potential of species i via EQUATION A.1 and enumerated the contributions
of the chemical potential and activity of species i in phase α via EQUATION A.2:

(A.1)

(A.2)

Furthermore, at equilibrium between the reactants i and products j, we have also established that ∆Grxn = 0 and that the electrochemical
potentials of the reactants and products are equal in the system:

(A.3)

Now, let us re-consider the Daniell cell example used in the main text (reproduced below in FIGURE A.1), where the individual half-reactions
(EQUATIONS A.4 and A.5) would be at equilibrium before electrical contact of the copper and zinc.

(A.4)

(A.5)

Despite the half-reactions being at equilibrium in their own compartments, a significant potential difference (over 1 V) exists between the metals.
If electrical connection is made between the two compartments, the following equilibrium reaction can be discussed:

(A.6)

Implied in EQUATION A.6 is the fact that zinc atoms in the zinc electrode lose electrons into the zinc electrode and electrons from the copper
metal reduce the copper cations. Thus, we can explicitly write this implication as:

(A.7)

When applying EQUATION A.3 to the above equilibrium reaction, it is helpful to specify the different phases in the Daniell cell. There are four
phases where reactants and products are present: the two solutions (here called phase α and phase β), as well as the solid Zn and Cu phases. These
phases are labeled in FIGURE A.1.

© 2023 American Chemical Society 12


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

FIGURE A.1 Daniell Cell with the Phases of Interest Indicated.

We can now expand EQUATION A.3 for all the species in EQUATION A.7 at equilibrium:

(A.8)

Rearranging, we find:

(A.9)

Additionally, given our previous discussion of electrochemical potential of electrons, the left side of the equation also can be expressed as:

(A.10)

where E is the potential difference between the Zn and Cu electrodes. Using EQUATION A.2, we can further expand EQUATION A.9 in terms of
standard chemical potentials, whereby:

(A.11)

becomes:

(A.12)

We now can relate the potential difference of our Daniell cell, E, to thermodynamic variables. Simply rearranging, we get:

(A.13)

© 2023 American Chemical Society 13


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

It is important to remember that this potential difference, E, is equivalent to the amount of work able to be done by passing a mole of unit charge
between the two electrodes. Therefore, it is inherently related to the change in Gibbs free energy, via:

(A.14)

where z is the number of moles of electrons passed per mole of reactant (e.g., z = 2 for EQUATION A.6) and F is the already defined Faraday’s
constant which converts z number of moles of electrons into coulombs of charge. A thermodynamic quantity, ΔG0 is defined as the standard
change in Gibbs free energy of a reaction and can similarly be expressed in terms of a potential difference: ΔG0 = −zFE0. This introduces the
concept of a standard cell potential, which is equivalent to the summation of the standard potentials for the reduction and oxidation
reactions giving rise to the potential difference as already introduced in SECTION 2.3.4. Mathematically, this can be expressed as:

(A.15)

In the case of EQUATION A.6, this can be expressed as:

(A.16)

Students of chemistry should also recognize that the standard change in Gibbs free energy should equal the difference in standard chemical
potentials of products and reactants. Thus, we can restate EQUATION A.13 as:

(A.17)

Dividing by −2F, gives:

(A.18)

This resembles the general form of an important relation in electrochemistry, the Nernst equation:

(A.19)

where Q is the reaction quotient. It is also clear that EQUATIONS A.18 and A.19 are not identical. EQUATION A.18 contains an additional term
(ϕα−ϕβ), which is the difference in potential between the two aqueous phases in this case. This type of potential difference is called a liquid
junction potential and occurs when any two dissimilar liquid phases come into contact. This alternative type of potential is discussed in SECTION
2.5.

© 2023 American Chemical Society 14


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

APPENDIX B
Derivation of the Butler–Volmer and Current-
Overpotential Equations
B.1 The Butler–Volmer Equation

B.2 The Current-Overpotential Equation

B.1 THE BUTLER–VOLMER EQUATION

Our goal is to quantify how the free energy of an electron-transfer reaction occurring at an electrode changes when we change the potential of
that electrode. Then, we will seek to use this to predict the current achieved at a given potential for a given reactant concentration at the electrode
surface.

To simplify our situation, let us consider the case in FIGURE 3.4 (copied below as FIGURE B.1) for a one-electron reduction reaction of oxidized
species O to reduced species R:

(B.1)

FIGURE B.1 Free Energy Diagram of Heterogeneous Electron-Transfer Reactions.

A free energy diagram for the heterogeneous, one-electron reduction reaction of O to R at two different applied potentials, E1 and
E2. The red trace (reaction 1) indicates the free energy surface of the reaction at the more negative potential (E1) and the blue trace
(reaction 2) indicates the free energy surface of the reaction at the more positive potential (E2). The free energy changes of activation
for either the cathodic reaction or the anodic reaction are labeled. The difference in free energy between the two
potentials is also labeled (FΔE).

Let us state the electrode in question is initially at equilibrium (at a potential E1 = E0′), but then the potential is changed to a new potential
(E2, where ΔE = E2 − E0′). The free energy of the reactants will change by FΔE. We cannot expect that the change in free energy of the activated
complex also changes by FΔE when we change the potential of the electrode by ΔE. Instead, the difference in free energy of the activated complex

© 2023 American Chemical Society 15


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

when using the two different potentials will be some fraction of this free energy, ((1 − α)FΔE), where α is termed the transfer coefficient and is a
value between 0 and 1. This coefficient is a measure of the symmetry between the reduced and oxidized energy surfaces and will not be discussed
further in this text.

Examining the free energies of activation for the forward and reverse reactions of EQUATION B.1, we find that:

(B.2)

and

(B.3)

Now that we have formulations for the change in the free energy of activation, we can report rate constants for the forward and reverse reactions,
assuming they follow the Arrhenius equation:

(B.4)

(B.5)

Inserting EQUATIONS B.2 and B.3 into EQUATIONS B.4 and B.5, respectively, gives us:

(B.6)

(B.7)

At equilibrium, the concentrations of O and R (cO* and cR*, respectively) will be equal. The * indicates that this concentration is a bulk
concentration; however, since this special case is at equilibrium, there is no net change in the concentration anywhere in solution (including at the
electrode’s interface). At equilibrium, the forward and reverse reactions are equal, i.e.:

(B.8)

Since the concentrations of O and R are also equal, this special case means that the forward and reverse rate constants are equal. We then define
the standard rate constant, k0, that is equal to these two equivalent rate constants that characterize equilibrium:

(B.9)

The forward and reverse rate constants can also be expressed in terms of the standard rate constant as at other potentials via:

(B.10)

(B.11)

Now we are ready to express the overall rate of reaction as a function of the rate constants derived and the concentrations at the electrode surface:

(B.12)

© 2023 American Chemical Society 16


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

We have acknowledged that the concentration of O or R at the surface can be time dependent. However, we can insert EQUATIONS B.10 and B.11
into EQUATION B.12 to get:

(B.13)

for any potential away from equilibrium by ΔE. In electrochemistry, the rate of reaction is innately related to the passage of current via:

(B.14)

with n being the number of electrons involved in the redox reaction and A being the surface area of the electrode that is in contact with the
redox solution. Thus, we arrive at a general kinetic expression of the current passing across the electrode surface that arises from the heterogeneous
electron-transfer reaction happening at the surface (assuming a one-electron process):

(B.15)

This expression is termed the Butler–Volmer equation (31, 32) and is an important equation in electrochemistry.

B.2 THE CURRENT-OVERPOTENTIAL EQUATION

One can perform a similar derivation under equilibrium conditions, where there is no net current. This means that the cathodic and anodic
reactions will have the same rate, such that:

(B.16)

where ΔE = Eeq − E0′, the Nernst equation applies, and the surface concentrations are equal to the bulk concentrations:

(B.17)

While EQUATION B.16 must be true and the cathodic and anodic currents must be equal at equilibrium, this does not mean that the currents need
to be zero. Thus, we can define some exchange current, i0, that is equal to either term in EQUATION B.16, i.e.:

(B.18)

Since the exchange current is defined at equilibrium, we can relate this current to the bulk concentrations of O and R by rearranging EQUATION
B.18:

(B.19)

Raising both sides of EQUATION B.19 to the power of −α yields:

(B.20)

which can now be substituted into EQUATION B.18, providing a way of expressing the exchange current as:

(B.21)

© 2023 American Chemical Society 17


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Furthermore, if we take the ratio of current (EQUATION B.15) to exchange current (EQUATION B.21) and solve for the current as a function of
exchange current, we get the current-overpotential equation (26):

(B.22)

where η is the overpotential—the difference between the potential applied to the working electrode and the equilibrium potential of the redox
reaction under study.

© 2023 American Chemical Society 18


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Bibliography
1. Simoska, O.; Minteer, S. D. Techniques in Electroanalytical Chemistry; American Chemical Society, 2021, 10.1021/acsinfocus.7e5021.

2. Fabbrizzi, L. Strange case of Signor Volta and Mister Nicholson: How electrochemistry developed as a consequence of an editorial misconduct. Angew. Chem., Int. Ed. 2019,
58 (18), 5810–5822, 10.1002/anie.201813519.

3. Galvani, L.; Aldini, G. Commentary on the Effect of Electricity on Muscular Motion; Elizabeth Licht, 1953.

4. Mertens, J. Shocks and sparks: The voltaic pile as a demonstration device. Isis 1998, 89 (2), 300–311, 10.1086/384002.

5. Wells, D. A. The Science of Common Things: A Familiar Explanation of the First Principles of Physical Science. For Schools, Families, and Young Students; Ivison, Phinney, Blakeman
& Co., 1859.

6. Ganot, A. Elementary Treatise on Physics: Experimental and Applied; William Wood & Co., 1893.

7. Volta, A.; Banks, J. I. On the electricity excited by the mere contact of conducting substances of different kinds. Philos. Mag. 1800, 7 (28), 289–311,
10.1080/14786440008562590.

8. Sarma, D. D.; Shukla, A. K. Building better batteries: A travel Back in time. ACS Energy Lett. 2018, 3 (11), 2841–2845, 10.1021/acsenergylett.8b01966.

9. Nicholson, W. Account of the electrical or galvanic apparatus of sig. Alex. Volta. J. Nat. Philos., Chem., Arts 1800, 4, 179–187.

10. Clarke, T. B.; Glasscott, M. W.; Dick, J. E. The role of oxygen in the voltaic pile. J. Chem. Educ. 2021, 98 (9), 2927–2936, 10.1021/acs.jchemed.1c00016.

11. Haldane, H. Experiments made with the metallic pile of signor Volta, principally directed to ascertain the powers of different metallic bodies. J. Nat. Philos., Chem., Arts 1800,
4, 313–319.

12. Davy, H. Notice of some observations on the causes of the galvanic phenomena, and on certain modes of increasing the powers of the galvanic pile of Volta. J. Nat. Philos.,
Chem., Arts 1800, 4, 337–342.

13. Silverstein, T. P. Oxidation and reduction: Too many definitions? J. Chem. Educ. 2011, 88 (3), 279–281, 10.1021/ed100777q.

14. Kragh, H. Confusion and controversy: Nineteenth-century theories of the voltaic pile. In Nuovo Voltiana; Bevilacqua, F.; Fregonese, L., Eds.; Vol. I; Università degli Studi di
Pavia, 2000; pp 133–157.

15. Kipnis, N. S. Changing a theory: The case of volts contact electricity. In Nuovo Voltiana; Bevilacqua, F.; Fregonese, L., Eds.; Vol. 5; Università degli Studi di Pavia, 2003; pp
143–162.

16. Ross, S. Faraday consults the scholars: The origins of the terms of electrochemistry. Notes Rec. R. Soc. Lond. 1961, 16 (2), 187–220, 10.1098/rsnr.1961.0038.

17. De Berg, K. C. The development of the theory of electrolytic dissociation. Sci. Educ. 2003, 12 (4), 397–419, 10.1023/A:1024438216974.

18. Scholz, F. Wilhelm Ostwald’s role in the genesis and evolution of the Nernst equation. J. Solid State Electrochem. 2017, 21 (7), 1847–1859, 10.1007/s10008-017-3619-y.

19. Falconer, I. J J Thomson and the discovery of the electron. Phys. Educ. 1997, 32 (4), 226–231, 10.1088/0031-9120/32/4/015.

20. Russell, C. A. The electrochemical theory of Sir Humphry Davy. Ann. Sci. 1959, 15 (1), 1–13, 10.1080/00033795900200018.

21. Russell, C. A. The electrochemical theory of Sir Humphry Davy. Ann. Sci. 1959, 15 (1), 15–25, 10.1080/00033795900200028.

22. Sacks, O. Everything in Its Place: First Loves and Last Tales; Vintage Books, 2019.

23. Cropper, W. H. James Joule's work in electrochemistry and the emergence of the first law of thermodynamics. Hist. Stud. Phys. Biol. Sci. 1988, 19 (1), 1–15,
10.2307/27757615.

24. Heard, D. M.; Lennox, A. J. J. Electrode materials in modern organic electrochemistry. Angew. Chem., Int. Ed. 2020, 59 (43), 18866–18884, 10.1002/anie.202005745.

25. Harris, D. C. Quantitative Chemical Analysis, Ninth Edition; W.H. Freeman and Co.: New York, NY, 2016.

26. Bakker, E.; Bühlmann, P.; Pretsch, E. Carrier-based ion-selective electrodes and bulk Optodes. 1. General characteristics. Chem. Rev. 1997, 97 (8), 3083–3132, 10.1021/
cr940394a.

27. Arons, A. B.; Bork, A. M. Newton's Laws of motion and the 17th century Laws of impact. Am. J. Phys. 1964, 32 (4), 313–317, 10.1119/1.1970268.

28. Varney, R. N.; Fisher, L. H. Electromotive force: Volta’s forgotten concept. Am. J. Phys. 1980, 48 (5), 405–408, 10.1119/1.12115.

© 2023 American Chemical Society 19


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

29. Saslow, W. M. Voltaic cells: The good (Faraday), the bad (Volta), and the ugly (Galvani). Phys. Teach. 2021, 59 (1), 22–26, 10.1119/10.0003010.

30. Bard, A. J.; Faulkner, L. R. Electrochemical Methods: Fundamentals and Applications; Wiley: New York, 2001.

31. Page, C. H. Electromotive force, potential difference, and voltage. Am. J. Phys. 1977, 45 (10), 978–980, 10.1119/1.10862.

32. Percival, S. J.; Bard, A. J. Ultra-sensitive potentiometric measurements of dilute redox molecule solutions and determination of sensitivity factors at platinum
Ultramicroelectrodes. Anal. Chem. 2017, 89 (18), 9843–9849, 10.1021/acs.analchem.7b01856.

33. Hamann, C. H.; Hamnett, A.; Vielstich, W. Electrochemistry; Wiley-VCH: Weinheim, 2007.

34. Zoski, C. G. Handbook of Electrochemistry; Elsevier: Amsterdam, Boston, 2007.

35. Butler, J. A. V. Studies in heterogeneous equilibria. Part II.—The kinetic interpretation of the Nernst theory of electromotive force. Trans. Faraday Soc. 1924, 19 (March),
729–733, 10.1039/TF9241900729.

36. Erdey-Grúz, T.; Volmer, M. Zur Theorie der Wasserstoff Überspannung. Z. Phys. Chem. 1930, 150A (1), 203–213, 10.1515/zpch-1930-15020.

37. Colburn, A. W.; Levey, K. J.; O'Hare, D.; Macpherson, J. V. Lifting the lid on the potentiostat: A beginner's guide to understanding electrochemical circuitry and practical
operation. Phys. Chem. Chem. Phys. 2021, 23 (14), 8100–8117, 10.1039/D1CP00661D.

38. Burstein, G. T. A hundred years of Tafel’s equation: 1905–2005. Corros. Sci. 2005, 47 (12), 2858–2870, 10.1016/j.corsci.2005.07.002.

39. Tafel, J. Über die Polarisation bei kathodischer Wasserstoffentwicklung. Z. Phys. Chem. 1905, 50U (1), 641–712, 10.1515/zpch-1905-5043.

40. Silverstein, T. P. Marcus theory: Thermodynamics CAN control the kinetics of Electron transfer reactions. J. Chem. Educ. 2012, 89 (9), 1159–1167, 10.1021/ed1007712.

41. Miller, J. R.; Calcaterra, L. T.; Closs, G. L. Intramolecular long-distance electron transfer in radical anions. The effects of free energy and solvent on the reaction rates. J. Am.
Chem. Soc. 1984, 106 (10), 3047–3049, 10.1021/ja00322a058.

42. Svante Arrhenius—Biographical. https://www.nobelprize.org/prizes/chemistry/1903/arrhenius/biographical/ (accessed November 15, 2022).

43. Arrhenius, S. Electrolytic dissociation. J. Am. Chem. Soc. 1912, 34 (4), 353–364, 10.1021/ja02205a001.

44. Masa, J.; Barwe, S.; Andronescu, C.; Schuhmann, W. On the theory of electrolytic dissociation, the greenhouse effect, and activation energy in (electro)catalysis: A tribute to
Svante Augustus Arrhenius. Chem. – Eur. J. 2019, 25 (1), 158–166, 10.1002/chem.201805264.

45. Arrhenius, S. Development of the Theory of Electrolytic Dissociation. In Nobel Lecture; 1903.

46. Myland, J. C.; Oldham, K. B. Uncompensated resistance. 1. The effect of cell geometry. Anal. Chem. 2000, 72 (17), 3972–3980, 10.1021/ac0001535.

47. Haynes, W. M.; Lide, D. R.; Bruno, T. J. CRC Handbook of Chemistry and Physics; CRC Press, 2016, 10.1201/9781315380476.

48. Pollok, D.; Waldvogel, S. R. Electro-organic synthesis—A 21st century technique. Chem. Sci. 2020, 11 (46), 12386–12400, 10.1039/D0SC01848A.

49. Frumkin, A.; Gorodetzkaja, A. Kapillarelektrische erscheinungen an amalgamen. Z. Phys. Chem. 1928, 136U (1), 451–472, 10.1515/zpch-1928-13634.

50. Frumkin, A. N.; Petrii, O. A.; Damaskin, B. B. Potentials of zero charge. In Comprehensive Treatise of Electrochemistry: The Double Layer; Bockris, J. O. M., Conway, B. E.,
Yeager, E., Eds.; Springer: US, 1980; pp 221–289.

51. Petrii, O. A. Adsorption phenomena on platinum group metal electrodes. Russ. Chem. Rev. 1975, 44 (11), 973, 10.1070/RC1975v044n11ABEH002403.

52. Bard, A. J.; Faulkner, L. F.; White, H. S. Electrochemical Methods: Fundamentals and Applications; Wiley, 2022.

53. Chapman, D. L. LI. A contribution to the theory of electrocapillarity. London, Edinburgh Dublin Philos. Mag. J. Sci. 1913, 25 (148), 475–481,
10.1080/14786440408634187.

54. Gouy, M. Sur la constitution de la charge électrique à la surface d'un électrolyte. J. Phys. Theor. Appl. 1910, 9 (1), 457–468, 10.1051/jphystap:019100090045700.

55. Vesterberg, O. History of electrophoretic methods. J. Chromatogr. A 1989, 480, 3–19, 10.1016/S0021-9673(01)84276-X.

56. Lemay, S. G.; Renault, C.; Dick, J. E. Particle mass transport in impact electrochemistry. Curr. Opin. Electrochem. 2023, 39, 101265, 10.1016/j.coelec.2023.101265.

57. Hickling, A. Studies in electrode polarisation. Part IV.—The automatic control of the potential of a working electrode. Trans. Faraday Soc. 1942, 38, 27–33, 10.1039/
TF9423800027.

58. Seh, Z. W.; Kibsgaard, J.; Dickens, C. F.; Chorkendorff, I.; Nørskov, J. K.; Jaramillo, T. F. Combining theory and experiment in electrocatalysis: Insights into materials design.
Science 2017, 355 (6321), eaad4998, 10.1126/science.aad4998.

© 2023 American Chemical Society 20


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

59. Fletcher, S. The theory of electron transfer. J. Solid State Electrochem. 2010, 14 (5), 705–739, 10.1007/s10008-009-0994-z.

60. Bard, A. J. Inner-sphere heterogeneous electrode reactions. Electrocatalysis and photocatalysis: The challenge. J. Am. Chem. Soc. 2010, 132 (22), 7559–7567, 10.1021/
ja101578m.

61. Elgrishi, N.; Rountree, K. J.; McCarthy, B. D.; Rountree, E. S.; Eisenhart, T. T.; Dempsey, J. L. A practical Beginner’s guide to cyclic voltammetry. J. Chem. Educ. 2018, 95
(2), 197–206, 10.1021/acs.jchemed.7b00361.

62. Chanda, D.; Xing, R.; Xu, T.; Liu, Q.; Luo, Y.; Liu, S.; Tufa, R. A.; Dolla, T. H.; Montini, T.; Sun, X. Electrochemical nitrogen reduction: Recent progress and prospects.
Chem. Commun. 2021, 57 (60), 7335–7349, 10.1039/D1CC01451J.

63. Morales-Guio, C. G.; Stern, L.-A.; Hu, X. Nanostructured hydrotreating catalysts for electrochemical hydrogen evolution. Chem. Soc. Rev. 2014, 43 (18), 6555–6569,
10.1039/C3CS60468C.

64. Fletcher, S. Tafel slopes from first principles. J. Solid State Electrochem. 2009, 13 (4), 537–549, 10.1007/s10008-008-0670-8.

65. Nørskov, J. K.; Bligaard, T.; Logadottir, A.; Kitchin, J. R.; Chen, J. G.; Pandelov, S.; Stimming, U. Trends in the exchange current for hydrogen evolution. J. Electrochem. Soc.
2005, 152 (3), J23, 10.1149/1.1856988.

66. Walczak, M. M.; Dryer, D. A.; Jacobson, D. D.; Foss, M. G.; Flynn, N. T. pH dependent redox couple: An illustration of the Nernst equation. J. Chem. Educ. 1997, 74 (10),
1195, 10.1021/ed074p1195.

67. Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J. R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for oxygen reduction at a fuel-cell cathode. J.
Phys. Chem. B 2004, 108 (46), 17886–17892, 10.1021/jp047349j.

68. Cheng, T.; Xiao, H.; Goddard, W. A., III. Reaction mechanisms for the electrochemical reduction of CO2 to CO and Formate on the Cu(100) surface at 298 K from
quantum mechanics free energy calculations with explicit water. J. Am. Chem. Soc. 2016, 138 (42), 13802–13805, 10.1021/jacs.6b08534.

© 2023 American Chemical Society 21


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Glossary
Activation energy: The minimum amount of energy required for a chemical reaction to proceed; the difference between the potential energy of
the reactants and the transition state (also known as the activated complex).

Adsorption isotherms: A relationship between the amount of adsorbed species as a function of pressure or concentration of species at constant
temperature.

Ag|AgCl: A reference electrode where the electrode is silver coated with silver (I) chloride (AgCl) and submerged in an aqueous solution of
potassium chloride. The potential is most commonly reported for saturated potassium chloride solutions at 0.197 V vs NHE; however, 3 M KCl
and 1 M KCl solutions are frequently used.

Amperometric experiment: An electroanalytical experiment where the applied potential at the working electrode is maintained at a constant
value during the experiment and the current is measured.

Anode: The electrode where oxidation occurs.

Arrhenius equation: A mathematical equation expressing the reaction rate constant (k) in terms of temperature (T), the activation energy of
the reaction (E a), and the universal gas constant (R).

Butler–Volmer equation: A kinetic equation that relates the current due to electrochemical reactions as a function of the potential difference
between the electrode and electrolyte solution.

Capacitance: The amount of charge stored per unit potential difference (having units of Farads, where 1 F = 1 C/V).

Capacitive current: Current that arises from charging or discharging the electric double layer (i.e., not from the transfer of electrons across the
interface to/from redox species.

Cathode: The electrode where reduction occurs.

Chemical potential: Formally, the change in free energy of an infinitely large system (e.g., a solution) per mole of “i” added (the partial
derivative of free energy with respect to change in mole fraction is expressed in the text).

Classical nucleation theory: A model to describe the process of forming a new phase that uses the changes in free energy during the growth
of a nanoparticle to predict the free energy barriers of phase formation, critical nucleus size, etc.

Conductance: A measure of the degree to which something conducts electricity; this value of an object is measured by dividing the current
running through the object by the applied potential across the object and is measured in Siemens (S) or inverse ohms (Ω−1).

Conductivity: An intrinsic property of a solution that measures the solution’s ability to conduct current; measured in Ω−1 cm−1 or S cm−1.

Convection: The mass transfer due to stirring or hydrodynamic transport.

Coulomb (C): Defined as 1 C = 1 A × 1 s, the coulomb is the SI unit of electric charge. It is equivalent to the charge of 6.241 × 1018 electrons,
as calculated from Faraday’s constant, F = 96,485 C/mol.

Counter or auxiliary electrode: One of the three electrodes used in the three-electrode cell, this electrode allows current to flow through the
cell without drawing substantial current through the reference electrode. The magnitude and direction of the redox reaction (i.e., oxidation or
reduction) occurring at the counter electrode will effectively be equal and opposite from the one occurring at the working electrode.

Critical radius: A measure of the smallest, stable size of a new phase beyond which the new phase is able to grow.

Current-overpotential equation: The kinetic equation that relates the current to the applied overpotential for the redox reaction.

Daniell cell: A historic electrochemical cell invented in 1836 by John Daniell where zinc was submerged in a zinc sulfate solution and copper
was submerged in a copper sulfate solution.

Density functional theory (DFT): A quantum mechanical simulation method used to calculate electronic structures and predict a wide range
of chemical features and phenomena.

Diffusion: The net movement of solute down its chemical potential gradient (e.g., from regions of higher concentration to regions of lower
concentration).

© 2023 American Chemical Society 22


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Diffusion coefficient: The factor of proportionality between the flux of a species and its concentration gradient. Larger diffusion coefficients
indicate faster the movement of the species by diffusion. The diffusion coefficient has unit of surface per time.

Dipole moment: A measure of the polarity or separation of charges within a molecule or system, having units of C m.

Double-layer capacitance: A measure of the ability for the electrode|solution to store electric charge as a function of potential.

Electric double layer: This structure is a manifestation of the contact between any two electrical conductors (it is especially relevant to
solid|solution interfaces). For example, a solid conductor surface not at 0 potential of zero charge will have some net charge and the first layer
of solution will accrue charge due to adsorbed ions, and the subsequent layers of solution will have differing degrees of ions due to Coulombic
interactions with the surface charge.

Electric potential energy: The energy held by a charge due to its position relative to all other charged particles in the system

Electrical mobility: The ability for a charged particle to move through a medium in the response of an electric field; also, the ratio of the drift
velocity of a charged species divided by the electric field strength causing the migration, expressed as μ, with units m2/(V·s).

Electrocatalysis: A type of catalysis where electrochemical reactions are driven at an electrode surface via the application of electrical potential.

Electrochemical potential: Formally, the change in free energy of an infinitely large system (e.g., a solution) per mole of charged species “i”
added that takes electrical effects from the species and environment into account.

Electrode: Any electronic conductor that contacts and can transfer current to a non-metallic medium, typically via reduction/oxidation
reactions. Examples include metals (e.g., Pt, Au), carbon (graphite), metallic liquid (Hg, amalgams), semiconductors (Si, indium–tin oxide).

Electrodeposition: The process by which electric current is used to deposit a material onto a conductive surface.

Electrolyte: Specifically refers to the ions dissolved in a solution. However, it is common in the electrochemistry literature (especially the energy
storage literature) for electrolyte to refer to any ionic conductor as the electrolyte (i.e., the entire solution that contains electrolytes).

Electrolytic cell or electrolysis cell: An electrochemical cell where an external voltage is applied to the two electrodes to drive a non-
spontaneous electrochemical reaction.

Electromotive force (emf): Often referring to a battery or galvanic cell, the maximum amount of work done per unit charge traveling
through an external circuit.

Electrostatic: Describing stationary electrical charges or forces.

Electrostatic force: The force between two electrically charged species at rest. The magnitude of this force is described by Coulomb’s law and
can be either attractive or repulsive depending on the sign of the charges.

Electrostatic potential: A position-dependent measure of the potential energy per unit charge. Defined as the amount of work required to
move a unit test charge from infinitely far away to that specific point.

Exchange current: An important parameter in the Butler–Volmer equation, this current is a measure of the amount of oxidative or reductive
current that is occurring when there is not net electrolysis occurring (at equilibrium). Just because there may be no net oxidation or reduction
occurring at an electrode, the rate of O being reduced to R is equal to the rate of R being oxidized to O. The exchange current is the current due
to this rate.

External circuit: Any combination of circuit elements and wires that connect the two electrodes in an electrochemical cell.

Faraday’s law: A law that equates the passage of 96,485.4 C of charge to the completion of 1 equivalent of reaction (for a one-electron
oxidation or reduction reaction). This gives us Faraday’s constant (96,485 C/mol). This allows for the integration of current with respect to time
to be associated with the amount of reaction performed.

Galvanic cell: An electrochemical cell that will spontaneously discharge upon the electrical connection between the two cell terminals as
chemical energy is converted to electricity.

Galvanostatic experiment: An electroanalytical experiment where the current is maintained at a constant value during the experiment and
the applied potential required to maintain this current is measured.

Gouy–Chapman theory: A model of the electric double layer wherein the surface charge on the electrode is not completely balanced by a layer
of excess (opposite) charge in solution; instead, this model assumes a diffuse layer of charge wherein the maximum amount of charge in solution is
closest to the surface layer and progressively lower net concentrations of ions extend into the solution.

© 2023 American Chemical Society 23


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Growth: In the context of phase transformation, the second step in the process by which a new phase begins to form, where the nuclei of the
new phase reach a critical size such that subsequent growth of the new phase is rapid and spontaneous.

Half-wave potential: The potential of an electrode at the point on a polarographic or voltammetric wave where the current is one half of the
diffusion current or the potential that is half-way between the oxidation and reduction peak of a cyclic voltamogram.

Inert: Not participating in (electro)chemical reactions. In reference to an electrode, allowing the transfer of electrons between the conductor and
the solution without itself being oxidized or reduced.

Inner-sphere reactions: Electron-transfer reactions that are characterized by a change in the bonding of the reactants or products at some
point during the electron-transfer process; a reaction where the redox species adsorbs/interacts with the electrode.

Inverted region: The domain of reactions where the rates of reaction (or rate constant) decreases as the reaction becomes more spontaneous
(the free energy of the reaction becomes more negative).

Kinetics: The study of chemical reaction rates.

Kohlrausch’s law: A relation characteristic of solutions with dilute concentrations of electrolyte that states that the molar conductivity linearly
increases as the square root of concentration decreases until a limiting conductivity is met. This model suggests that at infinite dilution, ions
migrate independently and the conductivity of a solution can be predicted by the sum of the molar conductivities.

Langmuir isotherm: A specific adsorption isotherm model that assumes adsorption is a reversible process and identical non-interacting
adsorption sites. See text for the equation that models this adsorption process.

Liquid junction potential: A potential difference between two dissimilar solutions. In many real electrochemical cells—even if the two
electrodes were initially placed in the same solution—upon the flow of current, the electrolyte solutions will become different near the electrodes.

Marcus theory: A theory originally proposed by Rudolph Marcus to explain the kinetics of outer-sphere reactions, wherein the rate of electron
transfer reactions is explained by assuming parabolic energy surfaces for both the reactants and products as a function of nuclear positions.

Mass transfer-limited: (In electrochemistry) the regime where the current is limited by the transport of species towards the electrode (rather
than the transfer of charge at the electrode).

Molar conductivity: The conductivity of an electrolyte solution per molar concentration, having units of S m2 mol−1.

Molecular dynamics: Computer simulations that predict how every atom and molecule in the system will behave over time based on a given
set of physical models for interatomic interactions.

Nernst diffusion layer thickness: A measure of the width that the non-uniform concentration profile of a redox species extends into the
bulk solution from the electrode.

Nernst equation: A mathematical expression relating potential to the concentrations of redox species at equilibrium in solution. This equation
is used to access the potential of individual electrodes, electrochemical cells, potential differences across liquid|liquid boundaries, amongst others.

Nernst–Planck equation: A differential equation that describes the flux of species in solution. This equation for flux has three terms to
quantify the flux due to three different types of mass transfer: diffusion, convection, and migration.

Non-Faradaic current: Current that is not due to the transfer of electrons across the interface to/from redox species (e.g., capacitive current).

Normal hydrogen electrode (NHE): A platinized Pt electrode in an acidic solution with 1 N proton and under 1 atm of molecular hydrogen
(H2); also accepted as the 0.0 V reference potential.

Nucleation: In the context of phase transformation, the first step in the process by which a new phase begins to form, where a cluster of atoms,
ions, or molecules forms that can be considered as a part of the new phase.

Ohmic drop: A potential difference that arises from the passage of current through a material due to it’s resistance. The ohmic drop can be
decreased by increasing the conductance (i.e., lowering the resistance of the solution) or decreasing the current.

Ohm’s law: Ohm’s law states that the current (i) is directly proportional to the potential difference (V) via the resistance (R). Written out, the
law state: V = i × R.

Outer-sphere reactions: Electron-transfer reactions that do not require nor cause any change in the bonding of the reactants or products;
a process by which the electrons can tunnel between the electrode and the redox species without interaction or adsorption of the species to the
electrode.

© 2023 American Chemical Society 24


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Overpotential: The difference between the applied potential at the working electrode and the equilibrium potential of the redox pair.

Oxidized: Having undergone oxidation, the process by which a species loses one or more electron(s). Historically, it was first used to describe the
combination with oxygen.

Physisorption: The physical process whereby a species is adsorbed to an interface with little perturbation of the electronic structure in the
species, primarily achieved via van der Waals and Coulombic interactions.

Poisson equation: Describing the relation between the divergence of electric displacement field and the density of mobile charges. In
electrochemistry, it’s solution can provide the potential in space arising from a system of charges.

Potential: A position-dependent measure of the potential energy per unit mass (gravitational potential) or per unit charge (electric potential).
Defined as the amount of work required to move a unit test mass (gravitational potential) or a unit test charge (electric potential) from infinitely
far away to that specific point.

Potential of zero charge, or pzc: The potential of an electrode wherein the surface has no excess of free charge (this is almost never 0 V vs a
given reference electrode).

Potential difference: In electrochemistry, the difference in electrochemical potential between two points. If measured between the two
electrodes in an electrochemical cell, this is the maximum energy to drive current in the external circuit between the electrodes and is the sum of
all individual potential differences in the electrochemical cell.

Potentiostat: An electronic instrument that is used to control voltage differences between electrodes (e.g., the working and reference electrode)
and measure currents (e.g., as a result of the applied potential to the working electrode).

Proton-coupled electron transfer reactions: An electrochemical reaction that also consumes or produces protons (H+) as a reactant or
product, respectively. These reactions are sensitive to pH changes.

Pseudocapacitance: A commonly misused term. See this link for its explanation.

Rate constants: The proportionality constant for a given chemical reaction between the rate of reaction and the concentration(s) of the
reactants; symbolized as k.

Reduced: Having undergone reduction, the process by which a species gains one or more electron(s).

Resistance: A measure of the degree to which something resists conducting electricity; this value of an object is measured in ohms (Ω).

Resistivity: A fundamental property of a material that measures its ability to resist the flow of current, having units of ohm meters (Ω m).

Salt bridge: A tube or other conduit containing an electrolyte (usually in a gel form) which can allow for electrical current to flow between two
solutions.

Saturated calomel electrode (SCE): A reference electrode where the electrode is mercury that is coated with calomel (mercury (I) chloride)
and submerged in an aqueous solution saturated with potassium chloride. The potential of this electrode is +0.244 V vs NHE.

Scan rate: The rate at which the applied potential is changed (having units of V/s or mV/s).

Specific adsorption: The chemical process whereby a species is adsorbed to an interface via a chemical interaction, analogous to a chemical
bond between reactants.

Standard hydrogen electrode (SHE): A platinized Pt electrode in a theoretically ideal acidic solution with unit activity of proton and
fugacity of molecular hydrogen (H2) at 1.00 bar; officially accepted as the 0.0 V reference potential, from which all other half-cell electrode
potentials are measured.

Standard rate constant: The rate constant of either the forward or reverse electron-transfer reactions under standard conditions at zero
overpotential (these will be equal under these conditions).

Standard reduction potential: Also known as standard electrode potentials, the standard reduction potential is the potential difference
between the measured electrode reaction and a platinum electrode submerged in a theoretically ideal acidic solution with unit activity of proton
and molecular hydrogen (H2) under standard conditions (298 K, 1 atm, and solutions with unit activity).

Stokes Law: A mathematical expression that expresses the frictional force exerted on a spherical particle in a fluid.

Strong electrolytes: These salts fully dissociate into solvated ions upon dissolution in water.

© 2023 American Chemical Society 25


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Supporting electrolyte: The electrolyte (dissolved ions in a solution) that serve to increase the conductivity of the solution, but do not
participate in the electrochemical reactions occurring during the experiment.

Surface tension: The work required to increase the surface area of a surface. The term “interfacial tension” is used when an interface is studied.

Tafel equation: An equation, first introduced in 1905, that relates the current with the overpotential.

Transfer coefficient: This coefficient (α) is a symmetry value between 0 and 1, reflecting the level of symmetry in the free energy surface
near the intersection between the reduced and oxidized energy surfaces on the reaction coordinate (i.e., the transition state for electrochemical
reactions). If the intersection is symmetrical, α = 0.5.

Transference number or transport number: The fraction of current flowing through an electrolyte solution due to a given ion, thereby
having values between 0 and 1. The sum of all transference numbers in an electrolyte solution will equal 1.

Transition state or activated complex: The highest potential energy state along the reaction coordinate of a chemical reaction; the
configuration of atoms during a chemical reaction that has the highest potential energy value.

Volcano plot: In electrochemistry, a scatterplot (having a shape that resembles a volcano) that often shows some value of the rate of
electrocatalytic reaction as a function of some measurement of binding energy (e.g., the free energy of adsorption).

Voltage (V): A measure of potential difference or electromotive force expressed in volts, the SI unit of voltage.

Voltage source: A system that can maintain a constant voltage between its two terminals, ideally independent of the current flowing through
it.

Voltaic pile: One of the world’s first recorded electric batteries; created by Alessandro Volta in 1800, it was a stack of repeating units of zinc,
brine-soaked cardboard, and silver (or copper).

Voltammetry: An electroanalytical experiment where a potential profile is applied at the working electrode over time and the current is
measured with respect to this time-dependent potential profile.

Weak electrolytes: These salts do not fully dissociate into solvated ions upon dissolution in water.

Working electrode: The electrode where the electrochemical reaction of interest is being measured (e.g., by applying a potential to this
electrode and measuring the current from the redox reaction of interest, measuring the potential of this electrode relative to a reference electrode,
etc.).

© 2023 American Chemical Society 26


https://pubs.acs.org/doi/book/10.1021/acsinfocus.7e7020

Footnotes

i All values taken from CRC Handbook of Chemistry and Physics, 97th ed.; CRC Press, 2016 (47). Dielectric constants reported at 20 °C.

© 2023 American Chemical Society 27

You might also like